• Aucun résultat trouvé

BUBBLING ABOVE THE THRESHOLD OF THE SCALAR CURVATURE IN DIMENSIONS FOUR AND FIVE

N/A
N/A
Protected

Academic year: 2021

Partager "BUBBLING ABOVE THE THRESHOLD OF THE SCALAR CURVATURE IN DIMENSIONS FOUR AND FIVE"

Copied!
42
0
0

Texte intégral

(1)

HAL Id: hal-01805092

https://hal.archives-ouvertes.fr/hal-01805092

Preprint submitted on 1 Jun 2018

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

BUBBLING ABOVE THE THRESHOLD OF THE

SCALAR CURVATURE IN DIMENSIONS FOUR

AND FIVE

Bruno Premoselli, Pierre-Damien Thizy

To cite this version:

Bruno Premoselli, Pierre-Damien Thizy. BUBBLING ABOVE THE THRESHOLD OF THE

(2)

CURVATURE IN DIMENSIONS FOUR AND FIVE.

BRUNO PREMOSELLI AND PIERRE-DAMIEN THIZY

Abstract. On any closed manifold (Mn, g) of dimension n ∈ {4, 5} we exhibit new blow-up configurations for perturbations of a purely critical stationary Schr¨odinger equation. We construct positive solutions which blow-up as the sum of two isolated bubbles, one of which concentrates at a point ξ where the potential k of the equation satisfies

k(ξ) > n − 2 4(n − 1)Sg(ξ),

where Sg is the scalar curvature of (Mn, g). The latter condition requires the bubbles

to blow-up at different speeds and forces us to work at an elevated precision. We take care of this by performing a construction which combines a priori asymptotic analysis methods with a Lyapounov-Schmidt reduction.

1. Introduction

1.1. Statement of the results. Let (M, g) be a smooth closed Riemannian manifold of dimension n ≥ 3. Let 4g = −divg(∇·) be the Laplace-Beltrami operator and let k be

a smooth function in M such that 4g+ k is coercive. We are interested in this paper

in the existence of energy-bounded blowing-up families of positive solutions (uε)ε>0 to

critical stationary Schr¨odinger equations of the following type:

4guε+ kεuε= u2

−1

ε in M, (1.1)

where 2∗ = n−22n is the critical power for the embedding of H1(M ) into Lebesgue spaces and (kε)ε>0 is a smooth perturbation of k. We say that a family (uε)ε>0 of solutions of

(1.1) has bounded energy if

lim sup

ε→0

kuεkH1(M )< +∞.

Since the work of Struwe [26] it is known that if (uε)ε>0 has bounded energy then, up to

a subsequence, there exist k ∈ N, k sequences (µ1,ε)ε, . . . , (µk,ε)ε of positive real numbers

converging to zero and k sequences (ξ1,ε)ε, . . . , (ξk,ε)ε of points of M such that

uε= u0+ k

X

i=1

Wi,ε+ o(1) in H1(M ), (1.2)

The first author is supported by a FNRS grant MIS F.4522.15.

(3)

where the Wi,ε are bubbling profiles given by Wi,ε=   µi,ε µ2 i,ε+ dg(ξi,ε,·)2 n(n−2)   n−2 2 , (1.3)

dg is the geodesic distance and uε * u0 as ε → 0. We say that the family (uε)ε>0 of

solutions of (1.1) blows-up if

lim sup

ε→0

kuεkC0(M ) = +∞.

If (uε)ε>0 has bounded energy and blows-up it is easily seen that k ≥ 1 in (1.2).

In the last decades, a vast amount of work was poured into understanding when equations (1.1) possess blowing-up families of positive solutions – with and without the

energy-bound assumption. It turns out that the geometric potential k ≡ cnSg plays a threshold

role, where we have let cn = 4(n−1)n−2 and where Sg denotes the scalar curvature of

(M, g). It was indeed proven in [6] that when n ≥ 4 (1.1) has no blowing-up positive solutions whatsoever if k < cnSg, and in [5] that (1.1) has no energy-bounded

blowing-up positive solutions if k > cnSg (unless maybe if n = 6, see [5]). When n = 3 the

situation is completely different, see [9]. The latter result is based on the generalization of decomposition (1.2) to C0(M ) obtained in [7]. On the other side, energy-bounded

blowing-up families of positive solutions of (1.1) have been constructed when kε is a

small perturbation of cnSg, see for instance [8, 17, 25]. In another direction, if kε≡ cnSg

for all ε, equation (1.1) is the Yamabe equation, and its compactness properties exhibit intriguing dimensional phenomena, see [2, 5, 11, 13, 14, 15].

In this article we construct, in dimensions four and five, exotic bubbling configurations for (1.1), where the potential k lies well above the threshold of the scalar curvature at one of the concentration points. In particular, our equations are not perturbations of the Yamabe equation. Let (M, g) be a n-dimensional closed Riemannian manifold, n ∈ {4, 5}, and let Ψ ∈ Cc∞(Rn) be a smooth compactly supported function in B0(R0) ⊂ Rn for

some R0 > 1. Assume that Ψ > 0 in B0(1) and that Ψ has a non-degenerate global

maximum at 0. Let ξ2,0∈ M be fixed. We let ig denote the injectivity radius of (M, g)

and for 0 < δ < ig/R0 we let hδ be given by

hδ(x) = Ψ  1 δexp −1 ξ2,0(x)  . (1.4)

In particular, hδ is supported in the geodesic ball Bg(ξ2,0, R0δ), and is allowed to change

sign if Ψ changes sign. Remember that (M, g) is said to be of positive Yamabe type if 4g+ cnSg is a positive operator. Our main result states as follows.

Theorem 1.1. Let (M, g) be a closed Riemannian manifold of dimension n ∈ {4, 5} of positive Yamabe type, not conformally diffeomorphic to the standard sphere (Sn, gstd).

Let ξ1,0 and ξ2,0 be distinct points in M and define hδ as in (1.4). Let δ > 0 small be

fixed and let H be any function in the class C(H) defined in (1.7) below. Then, for any 0 < ε ≤ ε0 small enough, there exists a positive solution uε of:

4guε+ cnSg+ hδ+ εHuε= u2

−1

(4)

in M . This family (uε)0<ε≤ε0 blows-up with finite energy at two distincts simple blow-up

points as ε → 0 and has a zero weak limit.

Note that when (M, g) is of positive Yamabe type and hδ is given by (1.4) the operator

4g+ cnSg+ hδ remains positive for small δ. On the other side, the nonnegativity of

4g+ cnSg+ hδ is a necessary condition to the existence of positive solutions of (1.5)

(see for instance [9], Lemma 2.1).

In the case of the standard sphere, remarkably, an analogue of Theorem 1.1 is available – unlike in the case of perturbations of the Yamabe equation when hδ ≡ 0 as investigated in

[8, 19]. Here the additional assumption that Ψ in (1.4) has negative average compensates for the vanishing of the Riemannian mass.

Theorem 1.2. Let ξ1,0 and ξ2,0 be distinct points in Sn, n ∈ {4, 5}, and define hδ as in

(1.4). Assume in addition that there holds: Z

R4

Ψ(y)dy < 0.

Let δ > 0 small be fixed and let H be any function in the class C(H) defined in (1.7) below. Then, for any 0 < ε ≤ ε0 small enough, there exists a positive solution uε of:

4guε+ n(n − 2) 4 + hδ+ εH  uε= u2 ∗−1 ε (1.6)

in M . This family (uε)0<ε≤ε0 blows-up with finite energy at two distincts simple blow-up

points as ε → 0 and has a zero weak limit.

The families (uε)ε of positive solutions that we construct in Theorems 1.1 and 1.2 blow

up as a sum of two isolated simple bubbles of nonequivalent weights. The highest one concentrates at a point ξ2,δ satisfying hδ(ξ2,δ) > 0, while the lowest one concentrates at

ξ1,0. In our constructions, δ is fixed small enough so that hδ(ξ1,0) = 0 by (1.4). To our

knowledge, Theorems 1.1 and 1.2 yield the first example of multi-bubble configurations in dimensions 4 and 5 when the limiting operator 4g+ cnSg+ hδ is positive; clustering

phenomena in the degenerate case had been previously constructed in [27, 29]. The weak limit of our families (uε)εis zero, and this is a necessary condition when n ∈ {4, 5}

by [5].

A few comments on the choice of hδ and H are in order here. First, we point out that

Theorems 1.1 and 1.2 require no smallness assumption on hδ(ξ2,0) = Ψ(0) and therefore

yield existence of (blowing-up) positive solutions for model equations like (1.1) when the limiting potential k is allowed to be much larger than cnSg at a blow-up point. Also, in

Theorem 1.2, no equivariance assumption is needed on hδ. The class C(H) of functions

H considered in Theorems 1.1 and 1.2 is defined as

C(H) = {H ∈ C∞(M ) satisfying (7.7) and (7.8) below }. (1.7) These functions H are used to construct the lowest bubble and can be chosen with great generality. By (1.7) we can choose H ≥ 0, in which case cnSg+ hδ+ εH in (1.5) (and

its counterpart in (1.6)) approaches cnSg+ hδ from above as ε → 0. Remark also that

since we assumed Ψ(0) > 0, the limiting potential always satisfies cnSg+ hδ> cnSg at

(5)

in Rn, which is possible in Theorem 1.1, it also satisfies cnSg+ hδ≥ cnSg everywhere in

M . However it does not satisfy cnSg+ hδ > cnSg everywhere in M , in adequation with

the results of [5]. Considering the additional hδ in the potential cnSg+ hδ brings in a

new set of technical problems which are not easily dealt with. They arise in Section 7, where the smallness assumption on δ is quantified and to which we refer for more details. We should also point out that, in some cases, Theorems 1.1 and 1.2 remain true when hδ given by (1.4) is replaced by a suitable smooth function h. Sufficient conditions on h

ensuring this are given in Remark 7.1 below.

Finally, the constructions that we produce here can only occur in dimensions 4 and 5. Indeed, as a consequence of the 3-dimensional sup-inf inequality, solutions of (1.1) can only blow-up as sums of bubbles of comparable weights when n = 3 (see [9], Theorem. 5.2, and see also [10] for other examples of bubbling phenomena in dimension 3). And when n ≥ 7, as a consequence of [5], energy-bounded families (uε)ε of solutions of (1.1)

only exist if limε→0kε = cnSg at all blow-up points (this remains true if n = 6 under

additional assumptions, see [9] prop. 8.1). Let us also mention that the picture when we drop the bounded-energy assumption is radically different: equivariant infinite-energy solutions when k > cnSg and when (Mn, g) is the standard sphere have been constructed

in [4] (when n ≥ 5) and very recently in [31] (when n = 4).

1.2. Strategy of proof of Theorems 1.1 and 1.2. First, we explain how an a priori blow-up analysis yields necessary conditions on the bubbling configuration in our setting. Assume that we are given a family (uε)ε of solutions of (1.5) (or (1.6)) that blows-up

with two bubbles – which are not a priori assumed to be isolated. By the H1-theory of [26] uε writes as

uε = W1,ε+ W2,ε+ o(1) in H1(M ), (1.8)

where Wi,ε, i = 1, 2 are given by (1.3) for some families (µi,ε)ε, i = 1, 2 of positive

numbers going to 0 and for some families (ξ1,ε)ε and (ξ2,ε)ε of points in M converging

towards ξ1 and ξ2 as ε → 0. If we assume now that h(ξ2) > 0, there is not much freedom

left: there necessarily holds that h(ξ1) = 0 and that (µ1,ε)ε and (µ2,ε)ε have to satisfy

µ1,ε= (C1+ o(1))µ2,εln  1 µ2,ε  if n = 4, µ31,ε= (C2+ o(1))µ2,ε if n = 5, (1.9)

as ε → 0, for positive constants C1, C2. In particular, ξ1 and ξ2 are distinct and ξ2,ε is

the center of the highest bubble. Similarly, the value of µ1,ε is constrained in terms of ε

by: µ1,εln  1 µ1,ε  = (C10 + o(1))ε if n = 4, µ1,ε= (C20 + o(1))ε if n = 5, (1.10)

for positive C10, C20. We refer to Appendix A where relations (1.9) and (1.10) are proven.

(6)

finite-dimensional reduction in H1(M ) to produce the constructions of Theorems 1.1 and 1.2 this would force us to work with an extremely high precision, since an expansion of Iε(W1,ε+ W2,ε) involves terms of order µ22,ε, where Iε is the energy functional of (1.5)

(or (1.6)). When n = 5, for instance, µ22,ε is comparable to µ61,ε in view of (1.9), which would force us to estimate the H1(M ) norm of the error in the nonlinear procedure with a precision o(µ31,ε). Finding a suitable ansatz for the approximated bubble W1,ε that

both reaches this precision and comes with explicit estimates to be able to compute the additional contributions in Iε(W1,ε+ W2,ε) seems both unnatural and technically out of

reach.

We overcome this technical difficulty by combining a priori pointwise asymptotic analysis techniques to a nonlinear finite-dimensional procedure in H1(M ). This new approach was recently developed by the first author in [21, 22] to construct instability examples for critical elliptic systems in a coupled supercritical setting. It goes as follows: we first perform the standard nonlinear procedure in H1(M ) and construct a candidate solution W1,ε+ W2,ε+ φε of (1.5) (or (1.6)) up to kernel elements, with φε controlled in

H1(M ). We use here the classical Lyapunov-Schmidt approach that has been developed in the last decades, see for instance [16, 20, 23, 25, 32] and the references therein. Since the H1(M ) bound on φε is not precise enough to proceed as usual, we then obtain a

thorough pointwise decription of the blow-up behavior of φε using techniques in the

spirit of those developed in [7] and [9]. In particular, we do not proceed via an expansion of the reduced-energy in our approach: we conclude our proof by showing that the kernel elements can be annihilated for suitable values of the parameters, and we use for this the latter pointwise estimates on φε.

In view of (1.9), the bubbling configurations that we investigate in this work can be thought of as the low-dimensional counterpart of towering phenomena in higher dimensions. Examples of towering phenomena for positive solutions have recently been constructed in dimensions n ≥ 7 in [17], carrying out a nice improvement of the usual energy methods, but taking advantage of a radial symmetry assumption. In this respect our approach, which relies on a priori analysis methods to perform the finite-dimensional reduction, allows us to overcome the absence of symmetry in the configuration of our bubbles. We believe our method will prove useful in future work when addressing the construction of involved bubbling configurations, for instance in the absence of symmetries.

The structure of the article is as follows. In Section 2 we introduce the bubbling profiles W1,ε and W2,ε. An elevated precision is required on W1,ε while a naive choice of W2,ε is

enough. In Section 3 we apply the standard nonlinear reduction procedure in H1(M ) and construct a solution W1,ε+ W2,ε+ φε of (1.5) up to kernel elements. Sections 4, 5 and 6

are the core of the analysis of the paper. In Section 4 we turn the H1 bound on φεinto a

global C0 one and show that φε = o(W1,ε+ W2,ε) in C0(M ). This requires an adaptation

of the techniques of [7], since W1,ε+ W2,ε+ φε is only a solution of (1.5) up to kernel

elements and can change sign. In Section 5 we improve the global estimate of Section 4 into a sharp higher-order pointwise control on φεaround ξ2. This again involves blow-up

(7)

elements. On one side, those pertaining to the kernel associated to the lowest bubble W1,ε are simply expanded using energy estimates. On the other side, those coming from

the highest bubble W2,ε cannot be dealt with in this way and are instead computed

using the precise pointwise asymptotics of Section 5. The analysis in Sections 4, 5 and 6 does not use (1.4) and can be performed in full generality. Section 7 contains the concluding vanishing argument in the proof of Theorems 1.1 and 1.2. Finally, Appendix A describes the a priori analysis considerations leading to (1.9).

Acknowledgments: The authors warmly thank Olivier Druet and Emmanuel Hebey for stimulating discussions and valuable comments on the manuscript.

2. Notations and bubbling profiles

Let (M, g) be a closed Riemannian manifold of dimension n ∈ {4, 5} of positive Yamabe type – that is, such that 4g+ cnSg is coercive, where cn= 4(n−1)n−2 and Sg is the scalar

curvature of (M, g). By the standard conformal normal coordinates result of Lee-Parker [12], there exists Λ ∈ C∞(M × M ) such that by letting Λξ= Λ(ξ, ·) there holds that:

Λξ(ξ) = 1, ∇Λξ(ξ) = 0, (2.1) that Sgξ(ξ) = 0, ∇Sgξ(ξ) = 0, 4gξSgξ(ξ) = 1 6|Wg(ξ)| 2 g, (2.2)

where Sgξ denotes the scalar curvature of the conformal metric gξ= Λ 4 n−2

ξ g, and that,

for any point ξ ∈ M there holds for arbitrarily large given N :  expgξξ ∗ gξ (y) = 1 + O(|y| N), (2.3)

C1-uniformly in ξ ∈ M and in y ∈ TξM , |y| ≤ C. Here exp gξ

ξ denotes the exponential

map for the metric gξ at ξ with the identification of TξM to Rnvia a smooth orthonormal

basis of TξM defined in an open set containing ξ. For any ξ ∈ M , we let Ggξ denote the

Green’s function of the operator 4gξ+ cnSgξ in M . Since n ∈ {4, 5}, the result of [12]

asserts that for any ξ ∈ M one has:

Ggξ(ξ, exp

ξ (y)) =

1 (n − 2)ωn−1

|y|2−n+ A(ξ) + O(|y|) (2.4)

as |y| → 0, where ωn−1 is the volume of the standard sphere Sn−1. The constant A(ξ) in

(2.4) is called the mass of Ggξ at ξ. It smoothly depends on ξ and there holds A(ξ) > 0

for any ξ ∈ M provided (M, g) is not conformally diffeomorphic to the standard sphere, and A ≡ 0 otherwise. For the sake of clarity we also recall the conformal covariance property of the conformal laplacian: for any v ∈ C∞(M ) and ξ ∈ M ,

 4g+ cnSg  (Λξv) = Λ2 ∗−1 ξ  4gξ+ cnSgξ  (v).

If Gg denotes the Green’s function of 4g+ cnSg in M this yields in particular that :

(8)

Let ξ1,0 and ξ2,0 be distinct points of M , and let r0 > 0 be such that

8r0 < min ig(M ), dg(ξ1,0, ξ2,0), inf

ξ∈Mdgξ(ξ1,0, ξ2,0), (2.6)

where ig denotes the injectivity radius of (M, g) and dg and dgξ respectively denote

the Riemannian distance associated to the metric g and gξ. Let H and h be smooth

functions in M . Assume that H is supported in Bgξ1,0(ξ1,0, 2r0), where Bgξ1,0 denotes

the geodesic ball with respect to the metric gξ1,0, that 4g+ cnSg+ h is coercive and that

h is supported in M \Bgξ1,0(ξ1,0, 2r0), so that the supports of h and H are disjoint. The

precise form of H and h will only come into play in Section 7, and we do not assume for now that (1.4) holds. The blow-up analysis performed in Sections 4 and 5 and the expansions in Section 6 will only rely on the assumption on their supports. Similarly, whether (M, g) is conformally diffeomorphic to the standard sphere or not only comes into play in Section 7.

Let µ1 > 0 and ξ1∈ M . Following [8] we define, for x ∈ M :

ˆ W1,µ1,ξ1(x) = (n − 2)ωn−1Ggξ1(ξ1, x)Λξ1(x)×          dgξ1(ξ1, x)n−2µ n−2 2 1  µ21+dgξ1(ξ1, x) 2 n(n − 2) 1−n2 if dgξ1(ξ1, x) < r0, rn−20 µ n−2 2 1  µ21+ r 2 0 n(n − 2) 1−n2 if dgξ1(ξ1, x) ≥ r0. (2.7)

For µ1 > 0 and ξ1∈ M , let T1,µ1,ξ1 be the unique solution in M of:



4g+ cnSg+ h



T1,t1,ξ1 = −h ˆW1,µ1,ξ1. (2.8)

It is a smooth function in M since h is supported in M \Bgξ1,0(ξ1,0, 2r0). Let χ ∈ C∞(R+)

be a smooth nonnegative function, with χ ≡ 1 in [0, r0] and χ ≡ 0 in [2r0, +∞). Define,

for µ1, µ2 > 0, ξ1, ξ2∈ M , and for x ∈ M :

W1,µ1,ξ1(x) = ˆW1,µ1,ξ1(x) + T1,µ1,ξ1(x), W2,µ2,ξ2(x) = χ(dg(ξ2, x))µ n−2 2 2 µ22+ dg(ξ2, x)2 n(n − 2) 1−n2 . (2.9)

As announced in the introduction, the choice of W2,µ2,ξ2 is rougher than the choice of

W1,µ1,ξ1; in particular, the conformal correction at ξ2 is not required. Note also that

(9)

kernel elements, for 1 ≤ j ≤ n and x ∈ M : Z1,0,µ1,ξ1(x) = (n − 2)ωn−1dgξ1(ξ1, x) n−2G gξ1(ξ1, x)χ  dgξ1(ξ1, x)  Λξ1(x) × µ n−2 2 1 dgξ1(ξ1, x)2 n(n − 2) − µ 2 1 ! µ21+dgξ1(ξ1, x) 2 n(n − 2) !−n2 , Z1,j,µ1,ξ1(x) = (n − 2)ωn−1dgξ1(ξ1, x) n−2G gξ1(ξ1, x)χ  dgξ1(ξ1, x)  Λξ1(x) × µ n 2 1   expgξξ1 1 −1 (x), ej(ξ1)  gξ1(ξ1) µ21+dgξ1(ξ1, x) 2 n(n − 2) !−n2 , Z2,0,µ2,ξ2 = χ (dg(ξ2, x)) µ n−2 2 2  dg(ξ2, x)2 n(n − 2) − µ 2 2   µ22+dg(ξ2, x) 2 n(n − 2) −n2 , Z2,j,µ2,ξ2 = χ (dg(ξ2, x)) µ n 2 2 D expξ2−1 (x), ej(ξ2) E gξ2(ξ2)  µ22+dg(ξ2, x) 2 n(n − 2) −n2 . (2.10) In (2.10) we denoted by the same notation (e1(y), · · · , en(y)) two families of orthonormal

vector fields, respectively for gξ1 and g, defined in open sets containing respectively ξ1

and ξ2.

We conclude this subsection with a remark. Let ξ1∈ Bgξ1,0(ξ1,0, r0). By (2.7), and since

h(y) = 0 for any dgξ1(ξ1, y) ≤ r0 by (2.6), T1,µ1,ξ1 in (2.8) is represented, with (2.1), (2.5)

and (2.7), as: T1,µ1,ξ1(x) = −(n − 2)ωn−1 n(n − 2) n−22 µ n−2 2 1 Z M

Gh(x, y)h(y)Gg(y, ξ1)dvg(y)

+O(µ

n+2 2

1 ),

(2.11)

where Gh denotes the Green’s function of 4g+ cnSg+ h in M and the O(µ

n+2 2

1 ) term is

in C2(M ) and is independent of the choice of µ1 and ξ1. Similarly we also obtain that,

for any y ∈ M \{ξ1}:

Gg(ξ1, x) = Gh(ξ1, x) +

Z

M

Gg(ξ1, y)h(y)Gh(y, x)dvg(y). (2.12)

The latter with and (2.7), (2.9) and (2.11) shows in particular that, for dgξ1(ξ1, x) ≥ r0,

we have: W1,µ1,ξ1 = (n − 2)ωn−1(n(n − 2)) n−2 2 µ n−2 2 1 Gh(ξ1, ·) + O(µ n+2 2 1 ) in C2(M ). (2.13) 3. Reduced problem in H1(M ) Let ε > 0 and let t1, t2 be positive numbers. We define:

(10)

As explained in the introduction, this choice of µ1,ε and µ2,ε is not a lucky guess but is

necessary and driven by conditions (1.9) and (1.10). The blowing-up solutions of (1.5) and (1.6) that we construct in this paper are bubbles modeled on (2.9) for the choice of µ1, µ2 given by (3.1). For t1, t2 > 0 and ξ1, ξ2 ∈ M we thus let, for 0 ≤ j ≤ n:

W1,ε,t1,ξ1 = W1,µ1,ε(t1),ξ1,

W2,ε,t1,t2,ξ2 = W2,µ2,ε(t1,t2),ξ2,

Z1,ε,j,t1,ξ1 = Z1,j,µ1,ε(t1),ξ1,

Z2,ε,j,t1,t2,ξ2 = Z2,µ2,ε(t1,t2),ξ2,

(3.2)

where µ1,ε(t1) and µ2,ε(t1, t2) are given by (3.1). Let A0 be a connected compact set in

(0, +∞), and let:

A = A0× Bgξ1,0(ξ1,0, r0) × A0× Bg(ξ2,0, r0). (3.3)

Throughout the paper, for the sake of clarity and since no confusion will occur, whenever (t1,ε, ξ1,ε, t2,ε, ξ2,ε)εwill denote a family of points in A, the families µ1,ε(t1,ε), µ2,ε(t1,ε, t2,ε),

W1,ε,t1,ε,ξ1,ε, Z1,j,ε,t1,ξ1,ε, W2,ε,t1,ε,t2,ε,ξ2,ε, Z2,j,ε,t1,ε,t2,ε,ξ2,ε, 0 ≤ j ≤ n will just be

de-noted by µ1, µ2, W1, . . . . Similarly, (t1,ε, ξ1,ε, t2,ε, ξ2,ε)ε will often simply be denoted by

(t1, ξ1, t2, ξ2). Given (t1,ε, ξ1,ε, t2,ε, ξ2,ε)ε, and adopting these notations, we will also let,

for any x ∈ M :

θ1(x) = µ1+ dgξ1(ξ1, x) and θ2(x) = µ2+ dg(ξ2, x). (3.4)

The points ξ1 and ξ2 will be thought of as the centers, respectively, of the lowest and

the highest bubble. By the choice of A in (3.3), they will always satisfy dgξ1,0(ξ1,0, ξ1) ≤ r0 and dg(ξ2,0, ξ2) ≤ r0,

so that by (2.6) the supports of Z1,j or H are disjoint from the supports of W2, Z2,k or

h, for 0 ≤ j, k ≤ n. In particular, W2 is supported in the region where h is nontrivial.

As a first step of our proof, we apply the standard finite-dimensional reduction scheme to this family of bubbling profiles. For any ε > 0 and (t1, ξ1, t2, ξ2) ∈ A, where A is as

in (3.3), let

Kε,t1,ξ1,t2,ξ2 = Span{Z1,j, Z2,k, 0 ≤ j, k ≤ n}, (3.5)

where the Zi,j are defined in (3.2), and let Kε,t11,t22 be its orthogonal for the scalar

product:

hu, vi = Z

M



h∇u, ∇vi + (cnSg+ h + εH)uv



dvg. (3.6)

In the following, all the H1(M )-norms appearing, denoted by k · kH1(M ), will be taken

with respect to this scalar product. Also, throughout this paper, if (fε)ε, (gε)ε denote

families of numbers or functions, the notation “fε . gε” will be used to denote the

existence of a positive constant C independent of ε such that fε ≤ Cgε for any ε small

enough. If gε≥ 0, we will also write “fε= O(gε)” to say that |fε| . gε.

(11)

Proposition 3.1. There exists ε0 > 0 such that for any 0 < ε ≤ ε0 and for any

(t1, ξ1, t2, ξ2) ∈ A, there exists φε(t1, ξ1, t2, ξ2) ∈ Kε,t⊥1,ξ1,t2,ξ2 such that

ΠK⊥ ε,t1,ξ1,t2,ξ2 " W1+ W2+ φε(t1, ξ1, t2, ξ2) −4g+ cnSg+ h + εH −1 W1+ W2+ φε(t1, ξ1, t2, ξ2) 2∗−1 + # = 0, (3.7) where W1and W2are as in (3.2) and where ΠK⊥

ε,t1,ξ1,t2,ξ2 denotes the orthogonal projection

on Kε,t

1,ξ1,t2,ξ2 for (3.6) . In addition, for any 0 < ε ≤ ε0, φε ∈ C

0(A, H1(M ))

and there exists a positive constant C such that, for any 0 < ε ≤ ε0 and for any

(t1,ε, ξ1,ε, t2,ε, ξ2,ε)ε∈ A there holds:

kφε(t1,ε, ξ1,ε, t2,ε, ξ2,ε)kH1(M ) ≤ Cεµ n−2

2

1 , (3.8)

where µ1 is given by (3.1) for t1= t1,ε. Also, φε(t1, ξ1, t2, ξ2) is the unique solution of

(3.7) in Kε,t

1,ξ1,t2,ξ2 satisfying (3.8).

In (3.7) we have let, for any u ∈ H1(M ), u+ = max(u, 0).

Proof. The existence, continuity and uniqueness properties of φε for 0 < ε ≤ ε0 for some

ε0 > 0, as well as (3.7), are a consequence of the general framework developed in [25]

(Proposition 5.1), in which (1.5) and (1.6) fall. The result of [25] generalizes previous ideas developed in [16, 20]. It remains to prove (3.8). Let (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε0 ∈ A. We

claim that the following estimate holds: there exists a positive constant C, independent on ε and on the choice of the family (t1,ε, ξ1,ε, t2,ε, ξ2,ε)ε, such that for any 0 < ε ≤ ε0,

W1+ W2−  4g+ cnSg+ h + εH −1 W1+ W2 2∗−1 H1(M ) ≤ Cεµ n−2 2 1 , (3.9)

where we used again the notations W1, W2, µ1, µ2, ξ1, ξ2 as above. First, a simple test

function computation using (3.2), together with Sobolev and trace inequalities shows that W1+ W2−  4g+ cnSg+ h + εH −1 W1+ W2 2∗−1 H1(M ) . (4g+ cnSg+ h + εH) (W1+ W2) − (W1+ W2) 2∗−1 Ln+22n (M ) + ||∂inW1+ ∂outW1|| L 2(n−1) n (∂B 1(ξ1,r0)) ,

where ∂inW1 and ∂outW1 denote the derivative with respect to the unit outward and

(12)

and H are disjoint we write, in M \∂Bgξ1(ξ1, r0), that: 4g+cnSg+ h + εH(W1+ W2) − (W1+ W2)2 ∗−1 = 4g+ cnSg+ εH ˆ W1− ˆW2 ∗−1 1 + εHT1+ ˆW2 ∗−1 1 − Wˆ1+ T1 2∗−1 + 4g+ cnSg+ hW2− W2 ∗−1 2 + W12∗−1+ W22∗−1− (W1+ W2)2 ∗−1 . (3.10) On one side, straightforward computations using (2.11) and (3.2) give:

εHT1+ ˆW 2∗−1 1 − Wˆ1+ T1 2∗−1 + W12∗−1+ W22∗−1− (W1+ W2)2 ∗−1 Ln+22n (M ) . εµ n−2 2 1 + µn−21 + (µ1µ2) n−2 2 .

On the other side, straightforward computations give that there holds:

|∂inW1+ ∂outW1| . µ

n+2 2

1 in ∂Bgξ1(ξ1, r0)

and that, both in Bgξ1(ξ1, r0) and in M \Bgξ1(ξ1, r0), there holds:

4g+ cnSg+ εH ˆ W1− ˆW2 ∗−1 1 .µ n+2 2 1 r n−4 1  µ21+ r 2 1 n(n − 2) −n2 , (3.11)

where we have let r1= dgξ1(ξ1, ·) (see for instance [8], Proposition 2.2). As a consequence:

4g+ cnSg+ εH  ˆ W1− ˆW2 ∗−1 1 Ln+22n (M )+||∂inW1+ ∂outW1||L2(n−1)n (∂B 1(ξ1,r0)) . εµ n−2 2 1 .

Finally, straightforward computations using (3.2) show that there holds, for any x ∈ M : 4g+ cnSg+ hW2(x) − W 2∗−1 2 (x) .µ n−2 2 2 θ2(x)2−n, (3.12)

where θ2 is defined in (3.4). This gives in the end:

4g+ cnSg+ hW2(x) − W 2∗−1 2 (x) Ln+22n (M ). µ n−2 2 2 .

Combining all these computations into (3.10) and using the explicit expression of µ1, µ2

given by (3.1) concludes the proof of (3.9). Estimate (3.8) then follows from (3.9) by

the result of [25]. 

4. C0-theory and uniform a priori pointwise estimates on φε

Let ε0 be given by Proposition 3.1. For 0 < ε ≤ ε0, let (t1, ξ1, t2, ξ2) ∈ A and let

φε= φε(t1, ξ1, t2, ξ2) be given by Proposition 3.1. Equation (3.7) shows that there exist

λεi,j = λεi,j(t1, ξ1, t2, ξ2), i = 1, 2, 0 ≤ j ≤ n, such that φε satisfies:

 4g+ cnSg+ h + εH  uε= (uε)2 ∗−1 + + X i,j λεi,j4g+ cnSg+ h + εH  Zi,j, (4.1)

where we have let:

(13)

and W1 = W1,ε,t1,ξ1, . . . are given by (3.2). Since φε ∈ K

ε,t1,ξ1,t2,ξ2, integrating (4.1)

against Zi,j for 1 ≤ i ≤ 2 and 0 ≤ j ≤ n and using (3.8) yields the existence of a positive

constant C such that for any 0 < ε ≤ ε0 and for any (t1,ε, ξ1,ε, t2,ε, ξ2,ε)ε≤ε0 ∈ A there

holds, for all 1 ≤ i ≤ 2 and 0 ≤ j ≤ n:

|λε

i,j| = |λεi,j(t1,ε, ξ1,ε, t2,ε, ξ2,ε)| ≤ Cεµ

n−2 2

1 . (4.3)

We aim at constructing a solution of (1.5) and (1.6) via (4.1) by finding an element (t1,ε, ξ1,ε, t2,ε, ξ2,ε)ε which annihilates all the λεi,j. This goes through an asymptotic

expansion in C0(A) of the λεi,j as ε → 0, where A is given in (3.3). However, as explained in the introduction, having h > 0 in the region where the center of the highest bubble W2 is expected to be localized requires these expansions to be carried out with a high

precision that cannot be reached with the mere H1 estimate (3.8). In this section we therefore obtain a priori global pointwise asymptotic estimates on φε. These will be

refined into sharp second-order estimates on φε around ξ2 in the next section.

We show that φε is, in a pointwise sense, globally small compared to W1 and W2:

Proposition 4.1. There exists ε1 > 0 and a family of positive numbers (νε)0<ε≤ε1 with

limε→0νε= 0 such that there holds, for any 0 < ε ≤ ε1 and for any (t1, ξ1, t2, ξ2) ∈ A:

φε(t1, ξ1, t2, ξ2)(x) ≤ νε  W1(x) + W2(x)  for any x ∈ M. (4.4)

Here again W1 and W2 are given by (3.2) and A is as in (3.3). In particular, up to

assuming that ε1 is small enough, we will assume that νε≤ 12 for 0 < ε ≤ ε1.

Proof of Proposition 4. The proof of Proposition 4 is divided into two Lemmas. The first one establishes, for a fixed ε, continuity properties of the mapping φε in strong

spaces.

Lemma 4.2. There exists ε1> 0 such that, for any 0 < ε ≤ ε1 the mappings:

(t1, ξ1, t2, ξ2) ∈ A 7−→ W1,ε,t1,ξ1 + W2,ε,t1,t2,ξ2 ∈ C

0(M )

(t1, ξ1, t2, ξ2) ∈ A 7−→ φε(t1, ξ1, t2, ξ2) ∈ C0(M )

are well-defined and continuous.

Proof. For the first map, the assertion simply follows from the explicit expression of the right-hand side given by (3.2) and by the regularity properties of Λξ1. We thus prove

the Lemma for the second map. First, by (3.1) and (3.8) we let ε1 > 0 be such that, for

any 0 < ε ≤ ε1 and any (t1, ξ1, t2, ξ2) ∈ A there holds:

kφε(t1, ξ1, t2, ξ2)kH1(M ) <

1 2K

−n−22

n , (4.5)

where we have let:

Kn= s 4 n(n − 2)ω 2 n n (4.6)

and ωn is the volume of the standard unit n-sphere. Let 0 < ε ≤ ε1 be fixed and

(14)

(4.1) and by an adaptation of Trudinger’s argument [30] (see also [9], Theorem 2.15) we get that

(W1+ W2+ φε)+∈ Ls(M )

for some s > 2∗. Then, with (4.1), a bootstrap procedure applies and shows that uε∈ C2(M ), and hence that φε∈ C0(M ). Note however that φε is not smooth on the

sphere {dgξ1,0(ξ1,0, y) = r0} since ˆW1 is not.

Let now (t1,k, ξ1,k, t2,k, ξ2,k)k≥1be a sequence of points of A converging towards (t1,0, ξ1,0, t2,0, ξ2,0)

and let φk = φε(t1,k, ξ1,k, t2,k, ξ2,k) for any k ≥ 1 and φ0 = φε(t1,0, ξ1,0, t2,0, ξ2,0). By

Proposition 3.1, φk→ φ0 in H1(M ) as k → +∞. Assume first that the sequence (φk)k

is uniformly (in k) bounded in L∞(M ). Then (3.2), (4.3) and standard elliptic theory in (4.1) show that every subsequence of (φk)k admits a subsequence which converges in

C0(M ), and therefore to φ0. In this case, thus, φk→ φ0 in C0(M ) as k → +∞.

We therefore assume that, up to a subsequence, kφkkL∞(M ) → +∞ as k → +∞. A

Green’s representation formula for uk given by (4.2) with (4.1) and standard properties

of Green’s functions (see [24]) show, since (W1 + W2+ φk)+ ≥ 0, that there exists a

positive constant Cε, independent of k, such that:

inf

M φk ≥ −Cε (4.7)

for any k ≥ 1. In particular, we might as well assume that maxM(φk)+ → +∞ as

k → +∞ and let xkbe such that φk(xk) = kφkkL∞(M )= maxMk)+→ +∞ as k → +∞.

We let µk= φk(xk)−

2

n−2 and, for any x ∈ B0(ig(M )/µk), we let gk = exp

xkg(µk·) and ˜ uk(x) = µ n−2 2 k uk(expxk(µkx)).

With (4.1), ˜uk satisfies, for any y ∈ B0(ig(M )/µk):

4gku˜k(y) + µ2k  cnSg+ h + εH  (yk)˜uk(y) = (˜uk(y))2 ∗−1 + + µ n+2 2 k X i,j λεi,j(t1,k, ξ1,k, t2,k, ξ2,k)  4g+ cnSg+ h + εH  Zi,j(yk),

where we have let yk = expxk(µky). By (4.7), by the definition of xk and since ε is fixed

throughout this proof there holds:

−Cεµ n−2 2 k ≤ ˜uk(y) ≤ 1 + C0µ n−2 2 k

for some positive constant C0 and for any y ∈ B0(ig(M )/µk). By (4.3) and standard

elliptic theory, ˜uk converges therefore in Cloc1 (Rn) to ˜u0, with 0 ≤ ˜u0 ≤ 1, solution of

4ξu˜0 = ˜u2

−1

0 .

By the definition of xk we also have ˜u0(0) = 1, so that the classification result in [3]

implies that k˜u0kL2∗(Rn)= K

−n−22

(15)

holds thus: Z Bxk(Rµk) |φk|2∗dvg ≥ Z Bxk(Rµk) |uk|2∗dvg+ O(µ n−2 2 k ) = Z B0(R) |˜uk|2 ∗ dvgk+ O(µ n−2 2 k ) = Z B0(R) |˜u0|2 ∗ dx + o(1) = (1 + εR)Kn−n+ o(1),

as k → +∞, where limR→+∞εR = 0. This is a contradiction with (4.5) for R and k

large enough and concludes the proof of the Claim. 

The second Lemma establishes a rough version of (4.4):

Lemma 4.3. Let ε1 be as in Lemma 4.2 and let (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε1 be a family of

points of A. Define, for any 0 < ε ≤ ε1:

νε:= φε(t1,ε, ξ1,ε, t2,ε, ξ2,ε) W1,ε,t1,ε,ξ1,ε+ W2,ε,t1,ε,t2,ε,ξ2,ε C0(M ) . Then νε→ 0 as ε → 0.

Proof. We prove Lemma 4.3 by contradiction, and therefore assume the existence of a sequence (εk)k, 0 < εk≤ ε1, with εk→ 0 as k → +∞, such that

φk W1+ W2 C0(M ) ≥ η0 (4.8)

for some η0 > 0, for all k ≥ 1. In (4.8), for the sake of simplicity and using the

previous notations, we simply wrote t1,εk = t1, ξ1,εk = ξ1, W1,εk,t1,εk,ξ1,εk = W1,

φεk(t1,k, ξ1,k, t2,k, ξ2,k) = φk and so on. We will keep these notations throughout the

proof of the Lemma and it will be implicit that we will be working with the quantities given by (3.2), associated to the sequences (εk)k and (t1,εk, ξ1,εk, t2,εk, ξ2,εk)k.

The proof of Lemma 4.3 consists in an asymptotic a priori analysis of the sequence (φk)k and is divided into several steps.

Step 1: local convergence. We first show that, for 1 ≤ i ≤ 2, there holds, up to a subsequence: µ n−2 2 i uk expξi(µi·) → U0 in C 1 loc(Rn), (4.9)

as k → +∞, where uk= uεk is as in (4.2), where µ1, µ2 are given by (3.1) and

U0(x) =  1 + |x| 2 n(n − 2) 1−n2 for x ∈ Rn. (4.10)

(16)

vi,k(x) = µ

n−2 2

i uk expξi(µix). Letting gi,k = exp

ξig(µi·), with (4.1) vi,k satisfies, for

any x ∈ B0(ig(M )/µi) and for xk = expξi(µix):

4gi,kvi,k(x) + µ 2 i cnSg+ h + εkH(xk)vi,k = (vi,k)2 ∗−1 + +X i,j λεk i,jµ n+2 2 i  4g+ cnSg+ h + εkH  Zi,j(xk). (4.11)

We show that vi,k is uniformly bounded in Cloc0 (Rn) by investigating its positive and

negative part separately. First, a straightforward adaptation of the arguments in [18] shows, with (4.11), that (vi,k)+= max(vi,k, 0) satisfies the following equation in a weak

sense:

4gi,k(vi,k)++ µ

2

i cnSg+ h + εkH(expξi(µi·))(vi,k)+ ≤ (vi,k)

2∗−1 + +X i,j λεk i,jµ n+2 2 i  4g+ cnSg+ h + εkH 

Zi,j(expξi(µi·))1vi,k>0.

(4.12) By (4.3) and (2.10) we have: X i,j λεk i,jµ n+2 2 i  4g+ cnSg+ h + εkH  Zi,j(xk) → 0 in Cloc0 (Rn) as k → +∞. Also, µ2i cnSg+ h + εkH 

expξi(µi·) → 0 in Cloc0 (Rn), and by the definition of vi,k, by

(3.8) and (4.2) there holds that:

lim

r→0lim supk→+∞

Z

Bx(r)

vi,k2∗(y)dy = 0 ∀x ∈ Rn.

Hence, an adaptation of Trudinger’s argument [30] to (4.12) shows that for any R > 0 there exists CR> 0 such that

k(vi,k)+kC0(B

0(R)) ≤ CR (4.13)

for k large enough.

Independently, let Gkdenote the Green’s function of 4g+ (cnSg+ h + εkH) in M and let

(xk)k be a sequence of points in M . By (3.2), (4.1) and (4.3) a representation formula

for uk gives:

uk(xk) & −εkµ

n−2 2

1 W1(xk) + W2(xk).

We used here that by (2.10) and (3.2) there exists a positive constant C depending only on n such that |Zi,j| ≤ CWi for i = 1, 2 and j = 0, . . . , n. Since by (2.6) and (3.3) there

holds lim infk→+∞dg(ξ1, ξ2) > 0, the latter inequality shows in particular that for any

x ∈ B0(ig(M )/µi):

vi,k & −εk− εkµn−21 µ

n−2 2

2 .

This shows that:

(vi,k)−→ 0 in Cloc0 (Rn) as k → +∞. (4.14)

Standard elliptic theory, with (4.13) and (4.14), shows with (4.11) that vi,k converges in

(17)

Step 2: uniform lower bound on φk. We now show that there exists a positive

sequence ηk→ 0 as k → +∞ such that, up to a subsequence,

φk(x) & −ηk W1+ W2(x) for any x ∈ M. (4.15)

Let (xk)k a sequence of points such that

φk W1+ W2 (xk) = inf x∈M φk W1+ W2 . (4.16)

Remember that W1+ W2 is positive in M . We write again a representation formula for

uk with (4.1) and (4.3), which gives:

 W1+ W2+ φk  (xk) & Z M Gk(xk, y)  W1+ W2+ φk 2∗−1 + dvg − εkµ n−2 2 1  W1+ W2  & −εkµ n−2 2 1  W1+ W2  + Z Bξ1(Rkµ1) Gε(xk, y)  W1+ W2+ φk 2∗−1 + dvg + Z Bξ2(Rkµ2) Gk(xk, y)  W1+ W2+ φk 2∗−1 + dvg, (4.17) where Rk > 0 is chosen so that Rkµi → 0 as k → +∞ for i = 1, 2 and such that

Bξ1(Rkµ1) and Bξ2(Rkµ2) are disjoint for all k. The integrals in (4.17) are estimated

with Fatou’s lemma and (4.9) which in turn, with (4.16), yields (4.15).

Step 3: Blow-up analysis. Step 2 shows in particular that, for k large enough,

uk= W1+ W2+ φk≥ 1 2 W1+ W2 in M. So uk actually solves, in M  4g+ cnSg+ h + εkH  uk= u2 ∗−1 k + X i,j λεk i,j  4g+ cnSg+ h + εkH  Zi,j.

Using (4.3), an adaptation of the blow-up analysis performed in [22] (Proposition 4.1, Steps 4, 5, 6), see also [28], shows that there holds, for any sequence (xk)k of points of

M : |φk(xk)| = o  W1(xk) + W2(xk)  .

Applying the latter to the sequence (xk)k that achieves the maximum point of W1φ+Wk 2

in M then yields a contradiction with (4.8), and concludes the proof of Lemma 4.3.  We now conclude the proof of Proposition 4.1. Let ε1 be as in Lemma 4.2. Then, again

by Lemma 4.2, for any 0 < ε ≤ ε1 there exists (t1,ε, ξ1,ε, t2,ε, ξ2,ε) ∈ A such that:

(18)

where A is as in (3.3). Let, for any 0 < ε ≤ ε1, νε be given by Lemma 4.3 for this

maximal family (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε1. Then, for any x ∈ M , for any 0 < ε ≤ ε1 and

for any (t1, ξ1, t2, ξ2) ∈ A, there holds by (4.18) that:

φε(t1, ξ1, t2, ξ2)(x) W1,ε,t1,ξ1(x) + W2,ε,t1,t2,ξ2(x) ≤ νε.

Since limε→0νε= 0, this proves (4.4) and concludes the proof of Proposition 4.1. 

The estimates on |φε(t1, ξ1, t2, ξ2))| given by Proposition 4.1 are, for a given ε, uniform

in the choice of (t1, ξ1, t2, ξ2). This is an important property of our analysis that will be

crucial in the final argument of the proof of Theorems 1.1 and 1.2.

A consequence of Proposition 4.1 is that for any 0 < ε ≤ ε1 and for any (t1, ξ1, t2, ξ2) ∈ A,

we now have

W1+ W2+ φε(t1, ξ1, t2, ξ2) ≥

1

2(W1+ W2) .

In particular, with (4.1), we now see that for any 0 < ε ≤ ε1 and for any (t1, ξ1, t2, ξ2),

uε given by (4.2) actually satisfies in M :

 4g+ cnSg+ h + εH  uε= u2 ∗−1 ε + X i,j λεi,j4g+ cnSg+ h + εH  Zi,j. (4.19)

5. Second-order pointwise estimates

In this section we refine the pointwise estimate on φε given by Proposition 4.1 in balls

of fixed radius centered at ξ2. These improved pointwise estimates will compensate for

the insufficient precision of (3.8) and will be the crucial ingredient of the asymptotic expansion of the λεi,j in Section 6. Let:

A1= A0× Bgξ1,0(ξ1,0, r0) × A0× Bg ξ2,0,

r0

2, (5.1)

where A0 is the compact set in (0, +∞) appearing in (3.3). The second-order estimates

that we obtain are as follows:

Proposition 5.1. There exists ε2 > 0 and C > 0 such that, for any 0 < ε ≤ ε2 and for

any family (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε2 ∈ A1, where A1 is defined in (5.1), we have:

• If n = 4, and for any x ∈ Bg(ξ2,ε,r20):

ε(x)| ≤ C νε µ1+ µ2 + µ2|ln θ2(x)| + µ1µ2 ln θ2(x) µ2  µ2 θ2(x)2 ! , (5.2)

• If n = 5, and for any x ∈ Bg(ξ2,ε,r20):

|φε(x)| ≤ C νε µ 3 2 1 + µ 3 2 2 + µ 3 2 2θ2(x) −1+ µ72 1 + µ 3 2 1| ln µ2|µ 3 2 2  µ2 θ2(x)2 32 ! . (5.3)

Here, as before, we have let φε= φε(t1,ε, ξ1,ε, t2,ε, ξ2,ε), µ1 and µ2 are given by (3.1), θ2

is as in (3.4) and νε is given by Proposition 4.1.

(19)

Lemma 5.2. Let ε1 and (νε)0<ε≤ε1 be given by Proposition 4.1 and let (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε1

be a family in A, where A is in (3.3). There exists 0 < ε2≤ ε1 and C > 0 such that:

• If n = 4, and for any x ∈ Bg(ξ2,ε, 2r0):

ε(x)| ≤ C νε µ1+ µ2 + µ2|ln θ2(x)| + µ1µ2 ln θ2(x) µ2  µ2 θ2(x)2 ! , (5.4)

• If n = 5, and for any x ∈ Bg(ξ2,ε, 2r0):

|φε(x)| ≤ C νε µ 3 2 1 + µ 3 2 2 + µ 3 2 2θ2(x)−1+ µ 7 2 1 + µ 3 2 1| ln µ2|µ 3 2 2  µ2 θ2(x)2 32 ! , (5.5)

where we used the same notations as in the statement of Proposition 5.1.

Proof. Let (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε1 ∈ A. Throughout the rest of this proof C will denote

a positive constant independent of ε, which might change from one line to another. We will adopt the same notations as before.

We first assume that n = 4. By (4.19) and since the supports of H and W2 are

disjoint, φε = φε(t1,ε, ξ1,ε, t2,ε, ξ2,ε) satisfies in M :  4g+1 6Sg+ h + εH   φε− X i,j λεi,jZi,j  =  W1+ W2+ φε 3 −W1+ W2 3 −   4g+ 1 6Sg+ h + εH  W1− W13  + 3W12W2+ 3W1W22 −   4g+1 6Sg+ h  W2− W23  . (5.6) For any 0 < ε ≤ ε1, let Gε be the Green’s function of 4g + 16Sg + h with Dirichlet

boundary condition on Bg(ξ2, 2r0) (remember that ξ2 = ξ2,ε). Let (xε)ε≤ε1 be a family

of points in Bg(ξ2, 2r0). For any 0 < ε ≤ ε1, if xε∈ Bg(ξ2, 2r0)\Bg(ξ2, r0), there holds

trivially:

ε(xε)| ≤ kφεkC0(2r

0\r0), (5.7)

where we have let kφεkC0(2r

0\r0) = kφεkC0(Bg(ξ2,2r0)\Bg(ξ2,r0)). Otherwise we write a

representation formula on Bg(ξ2, 2r0) for φεwith (5.6). This is possible since by definition

W1 is smooth in Bg(ξ2, 2r0) and therefore so is φε. Since the Z1,j and H are supported

outside of Bξ2(2r0) we get that:

(20)

where I1 = Z Bg(ξ2,2r0) Gε(xε, y)  W1+ W2+ φε 3 (y) −W1+ W2 3 (y) dvg(y), I2 = Z Bg(ξ2,2r0) Gε(xε, y)  4g+1 6Sg+ h  W1− W13 dvg(y), I3 = Z Bg(ξ2,2r0) Gε(xε, y) W12W2+ W1W22 dvg(y), I4 = Z Bg(ξ2,2r0) Gε(xε, y)  4g+1 6Sg+ h  W2− W23 dvg(y).

First, by (3.12) and standard properties of Green functions (see [24]) there holds:

I4≤ Cµ2|ln θ2(xε)| , (5.9)

where θ2 is as in (3.4). Then, by (3.2) and since by (2.6) there holds dgξ1(ξ1, y) > r0 for

any y ∈ Bξ2(2r0), it is easily seen that there holds:

I2 ≤ Cµ31. (5.10)

Straightforward computations using (3.2) show that there holds:

I3 ≤ C  µ2|ln θ2(xε)| + µ1µ2 ln θ2(xε) µ2  µ2 θ2(xε)2  . (5.11)

Finally, Proposition 4.1 shows that  W1+ W2+ φε 3 −W1+ W2 3 ≤ CW1+ W2 2 |φε|,

so that straightforward computations lead to:

I1≤ C  µ21+ µ2 ln θ2(xε) µ2  µ2 θ2(xε)2  kφεkC0(2r 0), (5.12)

where we have let kφεkC0(2r

0)= kφεkC0(Bξ2(2r0)). Combining (5.9)–(5.12) in (5.8) gives,

with (5.7) and (3.1), that there holds, for any (xε)0<ε≤ε1, xε∈ Bg(ξ2, 2r0):

|φε(xε)− 4 X j=0 λε2,jZ2,j(xε)| ≤ C kφεkC0(2r 0\r0)+ |λ ε 2,0|µ2+ 4 X j=1 |λε2,j|µ22+ µ2|ln θ2(xε)| +  µ21+ µ2 ln θ2(xε) µ2  µ2 θ2(xε)2  kφεkC0(2r 0)+ µ1µ2 ln θ2(xε) µ2  µ2 θ2(xε)2 ! . (5.13) By evaluating (5.13) at suitable points satisfying dg(xε, ξ2) ≤ µ2 one gets with (3.1) the

following estimate on the λε2,j, 0 ≤ j ≤ 4:

4

X

j=0

|λε2,j| ≤ Cµ2 kφεkC0(2r

(21)

so that, using Proposition 4.1, (5.13) improves into: |φε(xε)| ≤ C νε µ1+ µ2 + µ2|ln θ2(xε)| +  µ21+ µ2 ln θ2(xε) µ2  µ2 θ2(xε)2  kφεkC0(2r 0)+ µ1µ2 ln θ2(xε) µ2  µ2 θ2(xε)2 ! . (5.15) We now claim that the following result holds true:

Claim 5.3. There exists 0 < ε2 ≤ ε1 and C > 0 such that, for any 0 < ε ≤ ε2, there

holds:

εkC0(2r

0)≤ Cµ1. (5.16)

Proof. Remember that µ1 = µ1,ε(t1,ε) is given by (3.1). We proceed by contradiction

and assume that for some sequence (εk)k of positive numbers, εk→ 0 as k → +∞, there

holds

kkC0(2r

0) µ1 as k → +∞. (5.17)

As before, until the end of this Claim it will be implicit that all the quantities µ1, ξ1, . . .

depend on this subsequence (εk)k according to (3.1) and (3.2). We will let in particular

φk= φεk(t1, ξ1, t2, ξ2). Let (yk)k be a sequence of points such that |φk(yk)| = kφkkC0(2r0).

By (3.1) and (5.17) there holds

µ2|ln θ2(yk)| ≤ µ2| ln µ2| = o(kφkkC0(2r 0))

as k → +∞, and by Proposition 4.1 limk→+∞νεk = 0 so that (5.15) and (5.17) show

that

θ2(yk) ≤ Cµ2, (5.18)

where θ2 is defined in (3.4). For any y ∈ B0(µr02) we then let:

˜ φk(y) = 1 kφkkC0(2r 0) φk expξ2(µ2y). (5.19)

On one hand, (5.15) and (5.17) show that:

| ˜φk(y)| ≤ C ln(1 + |y|) (1 + |y|)2 + o(1) on B0 r0 µ2 . (5.20)

On the other hand, (3.2), (3.11), (3.12) and (5.17) show that:

µ22 kφkkC0(2r 0) −   4g+1 6Sg+ h  W1− W13  + 3W12W2+ 3W1W22 −   4g+ 1 6Sg+ h  W2− W23 !  expξ2(µ2·) 

(22)

˜ φ0 as k → +∞, where ˜φ0 solves in R4: 4euclφ˜0 = 3U02φ˜0+ 4 X j=0 ˜ λ02,jVj. (5.21)

Here eucl denotes the Euclidean metric, U0 is as in (4.10) and for any 0 ≤ j ≤ 4 we have

let: ˜ λ02,j = lim k→+∞ λεk 2,j µ2kφkkC0(2r 0) .

That this limit exists, up to a subsequence, is a consequence of (5.14) and (5.17). Also, in (5.21), Vj is defined in Rn for any n ≥ 3 by:

V0(y) =  |y|2 n(n − 2)− 1   1 + |y| 2 n(n − 2) −n2 , Vj(y) = yj  1 + |y| 2 n(n − 2) −n2 , for any 1 ≤ j ≤ n. (5.22)

Passing (5.20) to the limit shows that ˜φ0 ∈ L4(R4) so that integrating (5.21) against Vj

shows that ˜λ0

2,j = 0 for any 0 ≤ j ≤ 4. Now the result of [1] implies that:

˜

φ0 ∈ Vect{Vj, 0 ≤ j ≤ 4}.

To conclude the proof of Claim 5.3 we now prove that ˜φ0∈ Vect{Vj, 0 ≤ j ≤ 4}⊥, where

the orthogonal is taken for the usual scalar product in ˙H1(R4). This will imply that ˜

φ0 ≡ 0, which is a contradiction since | ˜φ0(˜y0)| = 1 by (5.18), where ˜y0 is the limit of 1

µ2 exp

−1

ξ2 (yk) as k → +∞.

To prove that ˜φ0∈ Vect{Vj, 0 ≤ j ≤ 4}⊥, we write that by Proposition 3.1 there holds,

for any 0 ≤ j ≤ 4:

hφk, Zji = 0,

where h·, ·i is the scalar product given by (3.6). Let now R > 0 be fixed. The latter equality implies, since Z2,j, 0 ≤ j ≤ n, is supported in Bg(ξ2, 2r0), that:

Z Bg(ξ2,Rµ2) h∇φk, ∇Z2,ji + 1 6Sg+ hφkZ2,j ! dvg= Z ∂Bg(ξ2,Rµ2) ∂νZ2,jφkdσg − Z Bg(ξ2,2r0)\Bg(ξ2,Rµ2)  4g+1 6Sg+ h  Z2,jφkdvg. (5.23)

Straightforward computations using (2.10) show that there holds, for y ∈ Bg(ξ2, 2r0):

4gZ2,0(y) − 3W2(y)2Z2,0(y)

≤ Cµ2θ2(y)−2,

4gZ2,j(y) − 3W2(y)2Z2,j(y)

≤ Cµ22θ2(y)−3, for 1 ≤ j ≤ 4,

(5.24)

where θ2 is as in (3.4). Now, (3.2) and (5.15) show that:

(23)

that Z Bg(ξ2,2r0)\Bg(ξ2,Rµ2) µ2θ2(·)−2φkdvg = o(µ2kφkkC0(2r 0)), and that Z Bg(ξ2,2r0)\Bg(ξ2,Rµ2) W22Z2,jφkdvg ≤  C (1 + R)3 + o(1)  µ2kφkkC0(2r 0), so that (5.23) becomes: Z Bg(ξ2,Rµ2) h∇φk, ∇Z2,ji + 1 6Sg+ hφkZ2,j ! dvg ≤  Cln(1 + R) 1 + R + o(1)  µ2kφkkC0(2r 0).

Dividing both sides by µ2kφkkC0(2r

0), using the definition of ˜φk in (5.19), letting first k

go to +∞ and then R → +∞ we obtain that:

Z

R4

h∇ ˜φ0, ∇Vjidx = 0.

This proves that ˜φ0 ∈ Vect{Vj, 0 ≤ j ≤ 4}⊥and, as explained above, gives a contradiction,

thus concluding the proof of Claim 5.3. 

Plugging (5.16) into (5.15) and using (3.1) concludes the proof of Lemma 5.2 for n = 4.

Assume now that n = 5. The proof is similar to the four-dimensional case. By (4.19) φε satisfies in M :  4g+ 3 16Sg+ h + εH   φε− X i,j λεi,jZi,j  =  W1+ W2+ φε 73 −W1+ W2 73 +W1+ W2 73 − W 7 3 1 − W 7 3 2 −   4g+ 3 16Sg+ h + εH  W1− W 7 3 1  −   4g+ 3 16Sg+ h  W2− W 7 3 2  . (5.25) For any 0 < ε ≤ ε1, let Gε be the Green’s function of 4g + 163Sg + h with Dirichlet

boundary condition on Bg(ξ2, 2r0). Let (xε)0<ε≤ε1 be any family of points in Bg(ξ2, 2r0).

For any 0 < ε ≤ ε1, if xε∈ Bg(ξ2, 2r0)\Bg(ξ2, r0), there holds trivially:

ε(xε)| ≤ kφεkC0(2r 0\r0).

Otherwise we write a representation formula for φε with (5.25). Since the Z1,j are

(24)

using (4.3), we get that there holds, for any (xε)0<ε≤ε1, xε ∈ Bg(ξ2, 2r0): |φε(xε) − 5 X j=0 λε2,jZ2,j(xε)| ≤ C kφεkC0(2r 0\r0)+ µ 7 2 1 + µ 3 2 2θ2(xε)−1+ µ 3 2 1  µ2 θ2(xε) 2 + µ21+  µ2 θ2(xε) 2! kφεkC0(2r 0) ! . (5.26) Evaluating again the latter estimate at suitable points satisfying dg(xε, ξ2,ε) ≤ µ2 one

gets, with (3.1), the following estimate:

5 X j=0 |λε 2,j| ≤ Cµ 3 2 2  kφεkC0(2r 0)+ µ 3 2 1  , (5.27)

so that, using (3.1) and Proposition 4.1, (5.26) improves into:

ε(xε)| ≤ C νε µ 3 2 1 + µ 3 2 2 + µ 7 2 1 + µ 3 2 2θ2(xε)−1+ µ 3 2 1  µ2 θ2(xε) 2 + µ21+  µ2 θ2(xε) 2! kφεkC0(2r 0) ! . (5.28)

As before, we prove the following claim:

Claim 5.4. There exists 0 < ε2 ≤ ε1 and C > 0 such that, for any 0 < ε ≤ ε2, there

holds: kφεkC0(2r 0)≤ Cµ 3 2 1. (5.29)

Proof. Here again we proceed by contradiction and assume that for some sequence (εk)k

of positive numbers, εk→ 0 as k → +∞, there holds

kφkkC0(2r 0)  µ

3 2

1 as k → +∞, (5.30)

using the same notations as in the proof of Claim 5.3. Let (yk)k be a sequence of points

such that |φk(yk)| = kφkkC0(2r

0). By (3.1), (3.4) and (5.30) there holds µ 3 2

2θ2(yk)−1 =

o(kφkkC0(2r

0)) so that (5.28) and (5.30) show that

θ2(yk) ≤ Cµ2. (5.31)

For any y ∈ B0(µr02) we let again:

˜ φk(y) = 1 kφkkC0(2r 0) φk expξ2(µ2y). (5.32)

Then (5.28) and (5.30) show that there holds:

| ˜φk(y)| ≤ C

1

(1 + |y|)2 + o(1) for any y ∈ B0

r0

µ2

(25)

As before, (3.11), (3.12), Proposition 4.1, (5.25), (5.30) and standard elliptic theory show that ˜φk converges, up to a subsequence, in Cloc1 (R5) towards ˜φ0 as k → +∞, where

˜ φ0 solves in R5: 4euclφ˜0= 7 3U 4 3 0φ˜0+ 5 X j=0 ˜ λ02,jVj, (5.34)

where eucl is the Euclidean metric, U0 is as in (4.10), the Vj are as in (5.22) and for any

0 ≤ j ≤ 5 we have let: ˜ λ02,j = lim k→+∞ λεk 2,j µ 3 2 2kφkkC0(2r 0) ,

which exists, up to a subsequence, by (5.27) and (5.30). Passing (5.33) to the limit shows that ˜φ0 ∈ L

10

3 (R5) so that integrating (5.34) against Vj shows first that ˜λ0

2,j = 0

for any 0 ≤ j ≤ 5 and then, by the result of [1], that: ˜

φ0 ∈ Vect{Vj, 0 ≤ j ≤ 5}. (5.35)

As before, we now prove that ˜φ0 ∈ Vect{Vj, 0 ≤ j ≤ 5}⊥. By Proposition 3.1 there holds

again

k, Zji = 0

for any 0 ≤ j ≤ 5, where h·, ·i is given by (3.6), so that for any R > 0 this implies that:

Z Bg(ξ2,Rµ2) h∇φk, ∇Z2,ji + 3 16Sg+ hφkZ2,j ! dvg= Z ∂Bg(ξ2,Rµ2) ∂νZ2,jφkdσg − Z Bg(ξ2,2r0)\Bg(ξ2,Rµ2)  4g+ 3 16Sg+ h  Z2,jφkdvg. (5.36)

Straightforward computations using (2.10) show that there holds, for y ∈ Bξ2(2r0):

4gZ2,0(y) − 7 3W2(y) 4 3Z2,0(y) ≤ Cµ 3 2 2θ2(y)−3, 4gZ2,j(y) − 7 3W2(y) 4 3Z2,j(y) ≤ Cµ 5 2 2θ2(y)−4, for 1 ≤ j ≤ 5, (5.37)

where θ2 is as in (3.4). Now, (3.2), (5.28) and (5.30) show that:

(26)

so that (5.36) becomes: Z Bg(ξ2,Rµ2) h∇φk, ∇Z2,ji + 3 16Sg+ hφkZ2,j ! dvg ≤  C 1 + R + o(1)  µ 3 2 2kφkkC0(2r 0).

Dividing both sides by µ

3 2

2kφkkC0(2r

0), letting first the k → +∞ and then R → +∞ gives

as before that ˜φ0 ∈ Vect{Vj, 0 ≤ j ≤ 5}⊥ and hence ˜φ0 ≡ 0. But this is a contradiction

with (5.35), since | ˜φ0(˜y0)| = 1 by (5.31), where ˜y0 is the limit of µ12 exp−1ξ2 (yk), and

concludes the proof of Claim 5.4. 

Now, plugging (5.29) into (5.28) yields:

|φε(xε)| ≤ C νε µ 3 2 1 + µ 3 2 2 + µ 7 2 1 + µ 3 2 2θ2(xε) −1 + µ 3 2 1  µ2 θ2(xε) 2! .

Writing down again a representation formula for (5.25) and using the latter to estimate the term involving φε then concludes the proof of Lemma 5.2 for n = 5. 

Note that the precision that we reach in Lemma 5.2 is related to the nature of φε, in

particular to the property of φε to be orthogonal to the kernel elements.

End of the proof of Proposition 5.1. Of course the constants ε2 and C given by Lemma

5.2 do depend on the choice of the family (t1,ε, ξ1,ε, t2,ε, ξ2,ε)ε≤ε1. To conclude the proof

of Proposition 5.1 we establish as before their uniformity. We only write the argument for n = 4 since the n = 5 case works identically. First, the right-hand side of (5.2) (seen as a continuous function in M ) is obviously continuous in (t1, ξ1, t2, ξ2). Therefore,

by Lemma 4.2, there exists a family (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε1 ∈ A such that for any

0 < ε ≤ ε1 there holds: φε(t1,ε, ξ1,ε, t2,ε, ξ2,ε) νε(µ1,ε+ µ2,ε) + µ2,ε|ln θ2(·)| + µ1,εµ2,ε ln θ 2,ε(·) µ2  µ2 θ2(·)2 C0(B g(ξ2,0,r0)) = sup (t1,ξ1,t2,ξ2)∈A φε(t1, ξ1, t2, ξ2) νε(µ1+ µ2) + µ2|ln θ2(·)| + µ1µ2 ln θ 2(·) µ2  µ2 θ2(·)2 C0(B g(ξ2,0,r0)) , (5.38) where ξ2,0 and r0 are as in (2.6), νε is given by Proposition 4.1 and A is given by (3.3).

Note in particular that ξ2,ε∈ Bg(ξ2,0, r0). Let ε2 and C be the constants associated to

this family (t1,ε, ξ1,ε, t2,ε, ξ2,ε)0<ε≤ε1 by Lemma 5.2. Let now ξ2 ∈ Bg(ξ2,0,

r0

2). Then:

Bg(ξ2,

r0

(27)

Therefore, for any 0 < ε ≤ ε2, sup (t1,ξ1,t2,ξ2)∈A1 φε(t1, ξ1, t2, ξ2) νε(µ1+ µ2) + µ2|ln θ2(·)| + µ1µ2 ln θ 2(·) µ2  µ2 θ2(·)2 C0(Bg 2,r02)) ≤ sup (t1,ξ1,t2,ξ2)∈A1 φε(t1, ξ1, t2, ξ2) νε(µ1+ µ2) + µ2|ln θ2(·)| + µ1µ2 ln θ 2(·) µ2  µ2 θ2(·)2 C0(B g(ξ2,0,r0)) ≤ C,

where in the last inequality we used that A1⊂ A and Bg(ξ2,0, r0) ⊂ Bg(ξ2,ε, 2r0) in order

to apply (5.38) and Lemma 5.2. This concludes the proof of (5.2).  6. Asymptotic Expansion along the kernel

For any 0 < ε ≤ ε2 and (t1, ξ1, t2, ξ2) ∈ A1 we let φε = φε(t1, ξ1, t2, ξ2) be given by

Proposition 3.1, where A1 is as in (5.1) and ε2 is given by Proposition 5.1. In this

Section we obtain an asymptotic expansion of the functions λεi,j, 1 ≤ i ≤ 2, 0 ≤ j ≤ n defined in (4.19). Throughout this section, all the asymptotic expansions that we will write hold in C0(A1).

6.1. Expansion of the λε1,j, 0 ≤ j ≤ n. We first obtain an asymptotic expansion of the λε1,j, 0 ≤ j ≤ n:

Lemma 6.1. The following expansions hold in C0(A1) as ε → 0, where A1 is as in

(5.1): If n = 4: k∇V0k2L2(R4)λε1,0(t1, ξ1, t2, ξ2) = 2C1(4)H(ξ1)t1− C2(4)A(ξ1) + C3(4) 2 Fh(ξ1)  e−2t1ε + o(e− 2t1 ε ), k∇Vjk2L2(R4)λε1,j(t1, ξ1, t2, ξ2) = 4  C1(4)∇jH(ξ1)t1− C2(4)∇jA(ξ1) + C3(4) ∇jFh(ξ1)  e−3t1ε + o(e−3t1ε ). (6.1) If n = 5: k∇V0k2L2(R5)λε1,0(t1, ξ1, t2, ξ2) = 4 3C1(5)H(ξ1)t 2 1− 2C2(5)t31A(ξ1) + 2C3(5)t31Fh(ξ1)  ε3+ o(ε3), k∇Vjk2L2(R5)λε1,j(t1, ξ1, t2, ξ2) = 5  C1(5)∇jH(ξ1)t31− C2(5)∇jA(ξ1)t41+ C3(5) ∇jFh(ξ1)t41  ε4+ o(ε4). (6.2)

In (6.1) and (6.2) the Vj are defined in (5.22), A(ξ1) denotes the mass of the Green’s

(28)

given by

Fh(ξ) =

Z

M

Gh(ξ, y)h(y)Gg(y, ξ1)dvg(y), (6.3)

C1(n), C2(n) are positive constants given by (6.13), and C3(n) is a positive constant

defined by (6.10) below.

The explicit values of the constants Ci(n) do not come into play in our final argument.

It is important to notice that the term A(ξ1) − Fh(ξ1) is just, in view of (2.12), the

mass of the Green’s function Gh(ξ1, ·) at ξ1 by analogy with (2.4). Note that the mass

of Gh(ξ1, ·) at ξ1 exists because h is supported in M \Bgξ1,0(ξ1,0, 2r0). If h were just a

smooth function in M the next order term in expansion (2.4) would likely be singular too, as explained in [24].

Proof. These expansions only require the H1 estimate on φε given by (3.8). By (3.2),

the Z1,j, 0 ≤ j ≤ n are supported in Bgξ1,0(ξ1,0, 2r0). Since W2, the Z2,j and h are

supported in M \Bgξ1,0(ξ1,0, 2r0) it is easily seen that (4.19) rewrites in Bgξ1,0(ξ1,0, 2r0)

as: n X j=0 λε1,j(t1, ξ1, t2, ξ2)  4g+ cnSg+ εH  Z1,j = εHT1 +  4g+ cnSg+ εH  ˆ W1+ φε − ˆW1+ φε 2∗−1 −h Wˆ1+ T1+ φε 2∗−1 − Wˆ1+ φε 2∗−1i , (6.4)

where T1 is given by (2.8) with µ1 given by (3.1). To estimate the λε1,j we integrate (6.4)

against Z1,j. First, by (2.11), (3.1) and (3.2) there holds, for 0 ≤ j ≤ n:

ε Z M HT1Z1,jdvg =              ( o(e−2t1ε ) if n = 4 o(ε3) if n = 5, if j = 0, ( o(e−3t1ε ) if n = 4 o(ε4) if n = 5, if 1 ≤ j ≤ n. (6.5)

Then, we write that Wˆ1+T1+ φε 2∗−1 − Wˆ1+ φε 2∗−1 − (2∗− 1) ˆW12∗−2T1 . ˆW1+ φε 2∗−3 |T1|2+ |T 1|2 ∗−1 + ( ˆW1+ φε)2 ∗−2 − ˆW12∗−2 |T1| . |φε|2 ∗−2 + ˆW12∗−3|φε||T1| + Wˆ2 ∗−3 1 + |φε|2 ∗−3 |T1|2+ |T1|2 ∗−1 . (6.6)

Straightforward computations with (2.11), (3.1) and (3.2) give that:

(29)

while (2.11), (3.1), (3.2) and (3.8) give, with H¨older’s inequality: Z M  |φε|2∗−2|T1| + ˆW12∗−3|φε||T1| + |φε|2∗−3|T1|2|Z1,j|dvg =              ( o(e−2t1ε ) if n = 4 o(ε3) if n = 5, if j = 0, ( o(e−3t1ε ) if n = 4 o(ε4) if n = 5, if 1 ≤ j ≤ n.

Combining the latter computations in (6.6) then gives:

− Z M h ˆ W1+ T1+ φε 2∗−1 − Wˆ1+ φε 2∗−1i Z1,jdvg = − (2∗− 1) Z M ˆ W12∗−2T1Z1,jdvg+              ( o(e−2t1ε ) if n = 4 o(ε3) if n = 5, if j = 0, ( o(e−3t1ε ) if n = 4 o(ε4) if n = 5, if 1 ≤ j ≤ n. (6.7)

With (2.11) and (3.2) it is now easily seen that there holds:

− (2∗− 1) Z M ˆ W12∗−2T1Z1,0dvg = C3(n)µn−21 Fh(ξ1) + o(µn−21 ) (6.8) and − (2∗− 1) Z M ˆ W12∗−2T1Z1,jdvg = nC3(n)µn−11 ∇jFh(ξ1) + o(µn−11 ), (6.9)

where µ1 is given by (3.1) and where C3(n) > 0 is given by the following expansion as

ε → 0: − Z M ˆ W12∗−1T1dvg = C3(n)µn−21 Fh(ξ1) + O(µn1), (6.10)

for Fh defined in (6.3). Finally, by (2.7), (3.1), (3.8) and since by construction φε is

orthogonal to the Z1,j, 0 ≤ j ≤ n, an adaptation of the computations in [8] (Section 6)

(30)

and Z M h 4g+ cnSg+ εH  ˆ W1+ φε − ˆW1+ φε 2∗−1i Z1,jdvg = 4  C1(4)∇jH(ξ1)t1− C2(4)∇jA(ξ1)  e−3t1ε + o(e− 3t1 ε ) if n = 4, 5C1(5)∇jH(ξ1)t13− C2(5)∇jA(ξ1)t41  ε4+ o(ε4) if n = 5, (6.12) where C1(n) and C2(n) are two positive constants defined by the following expansion as

ε → 0: 1 2 Z M |∇ ˆW1|2g+ (cnSg+ εH) ˆW12dvg− 1 2∗ Z M ˆ W12∗dvg = 1 nK −n n +       C1(4)H(ξ1)t1− C2(4)A(ξ1)  e−2t1ε + o(e−2t1ε ) if n = 4  C1(5)H(ξ1)t21− C2(5)A(ξ1)t31  ε3+ o(ε3) if n = 5. (6.13)

It remains to notice that by (3.2) there holds:

hZ1,j, Z1,ki = δjkk∇Vjk2L2(Rn)+ o(1)

for 0 ≤ j, k ≤ n, and even:

hZ1,0, Z1,ji =

(

o(e−t1ε) if n = 4

o(ε) if n = 5,

for 1 ≤ j ≤ n, in C0(A1) as ε → 0, where h·, ·i is given by (3.6) and Vj is as in (5.22).

With (3.1), (6.4), (6.5), (6.7), (6.8), (6.9), (6.11) and (6.12) this concludes the proof of

the Lemma. 

6.2. Expansion of the λε2,j, 0 ≤ j ≤ n. In this subsection we obtain an asymptotic expansion of the λε2,j. Unlike the case of the λε1,j, the expansion of the λε2,j now crucially relies on the precise pointwise asymptotics on φε obtained in Sections 4 and 5.

Lemma 6.2. The following expansions hold in C0(A1) as ε → 0, where A1 is as in

(31)

If n = 5: k∇V0k2L2(R5)λε2,0(t1, ξ1, t2, ξ2) = 4 3D1(5)h(ξ2)t 2 2− D2(5)(t1t2) 3 2Gh1, ξ2)  ε6+ o(ε6), k∇Vjk2L2(R5)λε2,j(t1, ξ1, t2, ξ2) = 5D1(5)∇jh(ξ2)t32− D2(5)t 3 2 1t 5 2 2∇jGh(ξ1, ξ2)  ε9+ o(ε9). (6.15)

In (6.14) and (6.15) the Vj are as in (5.22), Gh denotes the Green’s function of 4g+

cnSg+ h in M and its derivative is taken with respect to ξ2. Also, D1(n) and D2(n) are

positive constants given by (6.20) and (6.23) below.

Proof. As before, all the asymptotic expansions that we will write here take place in C0(A1). By (2.7), (2.8), (3.2) and (4.19) and since the Z1,j and H vanish on Bg(ξ2, 2r0),

for any 0 < ε ≤ ε2 and for any (t1, ξ1, t2, ξ2) ∈ A1 there holds:

n X j=0 λε2,j(t1, ξ1, t2, ξ2)  4g+ cnSg+ h  Z2,j =  4g+ cnSg ˆW1− ˆW2 ∗−1 1 + ˆW12∗−1− Wˆ1+ T1 2∗−1 +4g+ cnSg+ h  W2− W2 ∗−1 2 − (2∗− 1)W2∗−2 2 W1 −h(W1+ W2)2 ∗−1 − W12∗−1− W22∗−1− (2∗− 1)W22∗−2W1 i −h W1+ W2+ φε 2∗−1 − (W1+ W2)2 ∗−1 − (2∗− 1)(W1+ W2)2 ∗−2 φε i +  4g+ cnSg+ h  φε− (2∗− 1)W2 ∗−2 2 φε − (2∗− 1)h W1+ W2 2∗−2 − W22∗−2iφε. (6.16)

We integrate (6.16) against Z2,j for 0 ≤ j ≤ n. First, using (2.7), (2.11), (3.2) and (3.11)

we get that, for any 0 ≤ j ≤ n:

(32)

Mimicking the computations that led to (6.8) and (6.9) we get that: Z M "  4g+ cnSg+ h  W2− W2 ∗−1 2 # Z2,0dvg =    2D1(4)h(ξ2)εt1t22e− 2t1 ε + o(εe−2t1ε ) if n = 4 4 3D1(5)h(ξ2)ε 6t2 2+ o(ε6) if n = 5 , (6.18) and Z M "  4g+ cnSg+ h  W2− W2 ∗−1 2 # Z2,jdvg = ( 4D1(4)∇jh(ξ2)ε2t1t32e− 3t1 ε + o(ε2e−3t1ε ) if n = 4 5D1(5)∇jh(ξ2)ε9t32+ o(ε9) if n = 5 , (6.19)

where the positive constants D1(4) and D1(5) are defined by:

1 2 Z M |∇W2|2g+ (cnSg+ h)W22dvg− 1 2∗ Z M W22∗dvg = 1 nK −n n + ( D1(4)h(ξ2)εt1t22e− 2t1 ε + o(εe−2t1ε ) if n = 4 D1(5)h(ξ2)ε6t22+ o(ε6) if n = 5, (6.20)

where Kn is defined in (4.6). Similarly, direct computations using (2.13), (3.1) and (3.2)

show that − (2∗− 1) Z M W22∗−2W1Z2,0dvg =    − D2(4)εt2e− 2t1 ε Gh1, ξ2) + o(εe−2t1ε ) if n = 4, − D2(5)ε6(t1t2) 3 2Gh1, ξ2) + o(ε6) if n = 5, (6.21) and, for 0 ≤ j ≤ n, that

−(2∗−1) Z M W22∗−2W1Z2,jdvg =    − 4D2(4)ε2t22e−3t1ε ∇jGh1, ξ2) + o(ε2e−3t1ε ) if n = 4, − 5D2(5)ε9t 3 2 1t 5 2 2∇jGh(ξ1, ξ2) + o(ε9) if n = 5, (6.22) where ∇G(ξ1, ξ2) stands for the derivative of Gh(ξ1, ·) at ξ2. Also, in (6.21) and (6.22),

the positive constants D2(n) are given by the following expansion as ε → 0:

Z M W22∗−1W1dvg =    D2(4)εt2e− 2t1 ε Gh1, ξ2) + o(εe−2t1ε ) if n = 4 D2(5)ε6(t1t2) 3 2Gh1, ξ2) + o(ε6) if n = 5. (6.23)

(33)

and using (2.13) and (3.2) yields: Z M h (W1+ W2)2 ∗−1 − W12∗−1− W22∗−1− (2∗− 1)W22∗−2W1 i Z2,jdvg =              ( o(εe−2t1ε ) if n = 4 o(ε6) if n = 5, if j = 0, ( o(ε2e−3t1ε ) if n = 4 o(ε9) if n = 5, if 1 ≤ j ≤ n. (6.24)

We now estimate the components in (6.16) where φε appears:

Claim 6.3. There holds, as ε → 0, in C0(A1) :

Z M h W1+ W2+ φε 2∗−1 − (W1+ W2)2 ∗−1 − (2∗− 1)(W1+ W2)2 ∗−2 φε i Z2,jdvg =              ( o(εe−2t1ε ) if n = 4 o(ε6) if n = 5, if j = 0, ( o(ε2e−3t1ε ) if n = 4 o(ε9) if n = 5, if 1 ≤ j ≤ n. (6.25)

Proof. Let 0 < ε ≤ ε2. By Proposition 4.1 we can write that there holds, in M :

W1+ W2+ φε 2∗−1 − (W1+ W2)2 ∗−1 − (2∗− 1)(W1+ W2)2 ∗−2 φε . (W1+ W2)2 ∗−3ε|2. (6.26)

On one hand, using (3.1), (3.2), (4.4) and (6.26) we get that:

Références

Documents relatifs

Then Szabo [Sz] constructed pairs of isospectral compact manifolds with boundary that are not locally isometric, and Gordon and Wilson [GW3] generalized his construction to

The monotonicity statement involves suitable re- pulsivity properties of the rescaled Hamiltonian H λ in the focusing regime under the orthogonality condition (3.11) and the a

Raphael, On the orbital stability of the ground states and the singularity formation for the gravitational Vlasov Poisson system, preprint] a sharp criterion for the global existence

MARCUS, Large solutions of semilinear elliptic equations: existence, uniqueness and asymptotic behavior, Jl. MARCUS, Large solutions of semilinear elliptic equations

the case of symmetric spaces, isometry classes of such manifolds (N, g) are not determined by the Riemannian curvature tensor but we now need the curvature and the torsion tensor of

for positive solutions of scalar curvature equations in the region where the scalar.. curvature function is

One of us (Cantor [6]) has shown that the solvability of the Lichnerowicz-York equation for asymptotically flat maximal slices is equivalent to the given metric

(1) Identities or estimates relating the number and character of critical points (i.e., zeroes of the gradient) of solutions of elliptic equations with the boundary